首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Bacterial atrazine catabolism is initiated by the enzyme atrazine chlorohydrolase (AtzA) in Pseudomonas sp. strain ADP. Other triazine herbicides are metabolized by bacteria, but the enzymological basis of this is unclear. Here we begin to address this by investigating the catalytic activity of AtzA by using substrate analogs. Purified AtzA from Pseudomonas sp. strain ADP catalyzed the hydrolysis of an atrazine analog that was substituted at the chlorine substituent by fluorine. AtzA did not catalyze the hydrolysis of atrazine analogs containing the pseudohalide azido, methoxy, and cyano groups or thiomethyl and amino groups. Atrazine analogs with a chlorine substituent at carbon 2 and N-alkyl groups, ranging in size from methyl to t-butyl, all underwent dechlorination by AtzA. AtzA catalyzed hydrolytic dechlorination when one nitrogen substituent was alkylated and the other was a free amino group. However, when both amino groups were unalkylated, no reaction occurred. Cell extracts were prepared from five strains capable of atrazine dechlorination and known to contain atzA or closely homologous gene sequences: Pseudomonas sp. strain ADP, Rhizobium strain PATR, Alcaligenes strain SG1, Agrobacterium radiobacter J14a, and Ralstonia picketti D. All showed identical substrate specificity to purified AtzA from Pseudomonas sp. strain ADP. Cell extracts from Clavibacter michiganensis ATZ1, which also contains a gene homologous to atzA, were able to transform atrazine analogs containing pseudohalide and thiomethyl groups, in addition to the substrates used by AtzA from Pseudomonas sp. strain ADP. This suggests that either (i) another enzyme(s) is present which confers the broader substrate range or (ii) the AtzA itself has a broader substrate range.  相似文献   

2.
The effect of oxidative stress induced by neurotoxic metal ions on the properties of the brain 20S proteasome or multicatalytic proteinase complex (MPC) has been studied. Exposure of the 20S proteasome to increasing amounts of Fe(III), Fe(II), Cu(II) or Zn(II) affects its main hydrolytic activities: trypsin-like (T-L), chymotrypsin-like (ChT-L), peptidylglutamyl-peptide hydrolase (PGPH), branched-chain amino acid preferring (BrAAP) and caseinolytic activities, although in different ways. T-L activity showed gradual activation by both iron ions but inhibition by Cu(II) and Zn(II). ChT-L and PGPH activities were inhibited whereas BrAAP activity was widely activated by all the tested metal salts except for zinc ions. Moreover, the exposure to ferrous salt increased the degradation rate of casein. The functional effects appear to be linked to oxidation-induced modifications, as demonstrated by an increase of carbonyl groups following the exposure to metal ions. In addition, modifications induced by ferrous salt on the catalytic subunits were also supported by western blot analyses performed using anti-X, anti-Y and anti-Z antibodies. The results obtained clearly indicate that metal-catalyzed oxidation strongly affects the functions of the brain 20S proteasome, even though the catalytic subunits seem to be differently influenced by oxidative phenomena.  相似文献   

3.
Bacterial atrazine catabolism is initiated by the enzyme atrazine chlorohydrolase (AtzA) in Pseudomonas sp. strain ADP. Other triazine herbicides are metabolized by bacteria, but the enzymological basis of this is unclear. Here we begin to address this by investigating the catalytic activity of AtzA by using substrate analogs. Purified AtzA from Pseudomonas sp. strain ADP catalyzed the hydrolysis of an atrazine analog that was substituted at the chlorine substituent by fluorine. AtzA did not catalyze the hydrolysis of atrazine analogs containing the pseudohalide azido, methoxy, and cyano groups or thiomethyl and amino groups. Atrazine analogs with a chlorine substituent at carbon 2 and N-alkyl groups, ranging in size from methyl to t-butyl, all underwent dechlorination by AtzA. AtzA catalyzed hydrolytic dechlorination when one nitrogen substituent was alkylated and the other was a free amino group. However, when both amino groups were unalkylated, no reaction occurred. Cell extracts were prepared from five strains capable of atrazine dechlorination and known to contain atzA or closely homologous gene sequences: Pseudomonas sp. strain ADP, Rhizobium strain PATR, Alcaligenes strain SG1, Agrobacterium radiobacter J14a, and Ralstonia picketti D. All showed identical substrate specificity to purified AtzA from Pseudomonas sp. strain ADP. Cell extracts from Clavibacter michiganensis ATZ1, which also contains a gene homologous to atzA, were able to transform atrazine analogs containing pseudohalide and thiomethyl groups, in addition to the substrates used by AtzA from Pseudomonas sp. strain ADP. This suggests that either (i) another enzyme(s) is present which confers the broader substrate range or (ii) the AtzA itself has a broader substrate range.  相似文献   

4.
Histone deacetylase 8 (HDAC8) was originally classified as a Zn(II)-dependent deacetylase on the basis of Zn(II)-dependent HDAC8 activity in vitro and illumination of a Zn(II) bound to the active site. However, in vitro measurements demonstrated that HDAC8 has higher activity with a bound Fe(II) than Zn(II), although Fe(II)-HDAC8 rapidly loses activity under aerobic conditions. These data suggest that in the cell HDAC8 could be activated by either Zn(II) or Fe(II). Here we detail the kinetics, thermodynamics, and selectivity of Zn(II) and Fe(II) binding to HDAC8. To this end, we have developed a fluorescence anisotropy assay using fluorescein-labeled suberoylanilide hydroxamic acid (fl-SAHA). fl-SAHA binds specifically to metal-bound HDAC8 with affinities comparable to SAHA. To measure the metal affinity of HDAC, metal binding was coupled to fl-SAHA and assayed from the observed change in anisotropy. The metal KD values for HDAC8 are significantly different, ranging from picomolar to micromolar for Zn(II) and Fe(II), respectively. Unexpectedly, the Fe(II) and Zn(II) dissociation rate constants from HDAC8 are comparable, koff ∼0.0006 s−1, suggesting that the apparent association rate constant for Fe(II) is slow (∼3 × 103 M−1 s−1). Furthermore, monovalent cations (K+ or Na+) that bind to HDAC8 decrease the dissociation rate constant of Zn(II) by ≥100-fold for K+ and ≥10-fold for Na+, suggesting a possible mechanism for regulating metal exchange in vivo. The HDAC8 metal affinities are comparable to the readily exchangeable Zn(II) and Fe(II) concentrations in cells, consistent with either or both metal cofactors activating HDAC8.  相似文献   

5.
Human tyrosine 3-monooxygenase (tyrosine hydroxylase) exists as four different isozymes (TH1-TH4), generated by alternative splicing of pre-mRNA. Recombinant TH1, TH2 and TH4 were expressed in high yield in Escherichia coli. The purified isozymes revealed high catalytic activity [when reconstituted with Fe(II)] and stability at neutral pH. The isozymes as isolated contained 0.04-0.1 atom iron and 0.02-0.06 atom zinc/enzyme subunit. All three isozymes were rapidly activated (13-40-fold) by incubation with Fe(II) salts (concentration of iron at half-maximal activation = 6-14 microM), and were inhibited by other divalent metal ions, e.g. Zn(II), Co(II) and Ni(II). They all bind stoichiometric amounts of Fe(II) and Zn(II) with high affinity (Kd = 0.2-3 microM at pH 5.4-6.5). Similar time courses were observed for binding of Fe(II) and enzyme activation. In the absence of any free Fe(II) or Zn(II), the metal ions were released from the reconstituted isozymes. The dissociation was favoured by acidic pH, as well as by the presence of metal chelators and dithiothreitol. The potency of metal chelators to remove iron from the hydroxylase correlated with their ability to inhibit the enzyme activity. These studies show that tyrosine hydroxylase binds iron reversibly and that its catalytic activity is strictly dependent on the presence of this metal.  相似文献   

6.
Ultraviolet difference spectroscopy has been used to study Fe (III)-apoferritin complexes formed after addition of Fe (II) to apoferritin in air. At constant iron, the recorded spectra varied with time after Fe (II) addition and with the number of iron atoms/molecule (protein concentration). The results indicate that after production of an initial complex, rearrangement or migration of Fe (III) atoms occurs, with polynuclear species forming as end-product, probably by hydrolytic polymerization. The presence of Tb3+ or Zn2+ ions affected the Fe (III) spectra and their development in different ways. The combined data suggest that more than one site, or processes, are involved in ferritin iron-core formation and that some of the metal sites are clustered.  相似文献   

7.
The gene encoding for the methionyl aminopeptidase from the hyperthermophilic archaeon Pyrococcus furiosus (PfMetAP-II; EC 3.4.11.18) has been inserted into a pET 27b(+) vector and overexpressed in Escherichia coli. The new expression system resulted in a 5-fold increase in purified enzyme obtained from a 5 L fermentor growth. The as-purified PfMetAP-II enzyme, to which no exogenous metal ions or EDTA was added, was found to have 1.2 equiv of zinc and 0.1 equiv of iron present by ICP-AES analysis. This enzyme had a specific activity of 5 units/mg, a 60-fold decrease from the fully loaded Fe(II) enzymes. When an additional 2 equiv of Zn(II) was added to the as-purified PfMetAP-II, no activity could be detected. The combination of these data with previously reported whole cell studies on EcMetAP-I further supports the suggestion that the in vivo metal ion for all MetAP's is Fe(II). Both Co(II)- and Fe(II)-loaded PfMetAP-II showed similar substrate specificities to EcMetAP-I. Substrate binding was largely affected by the amino acid in the P1 position and the length of the polypeptide. The substrates MSSHRWDW and MP-p-NA showed the smallest K(m) values while the substrates MGMM and MP-p-NA provided the highest turnover. The catalytic efficiency (k(cat)/K(m)) of PfMetAP-II for MP-p-NA at 30 degrees C was 799 500 and 340 930 M(-1) s(-1) for Co(II)- and Fe(II)-loaded PfMetAP-II, respectively. Maximum catalytic activity was obtained with 1 equiv of Co(II) or Fe(II), and the dissociation constants (K(d)) for the first metal binding site were found to be 50 +/- 15 and 20 +/- 15 nM for Co(II)- and Fe(II)-substituted PfMetAP-II, respectively. Electronic absorption spectral titration of a 1 mM sample of apo-PfMetAP-II with Co(II) provided a dissociation constant of 0.35 +/- 0.02 mM for the second metal binding site, a 17500-fold increase compared to the first metal binding site. The electronic absorption data also indicated that both Co(II) ions reside in a pentacoordinate geometry. PfMetAP-II shows unique thermostability and the optimal temperature for substrate turnover was found to be approximately 85 degrees C at pH 7.5 in 25 mM Hepes and 150 mM KCl buffer. The hydrolysis of MGMM was measured in triplicate between 25 and 85 degrees C at eight substrate concentrations ranging from 2 to 20 mM. Both specific activity and K(m) values increased with increasing temperature. An Arrhenius plot was constructed from the k(cat) values and was found to be linear over the temperature range 25-85 degrees C, indicating that the rate-limiting step in PfMetAP-II peptide hydrolysis does not change as a function of temperature. Co(II)- and Fe(II)-loaded PfMetAP-II have similar activation energies (13.3 and 19.4 kJ/mol, respectively). The thermodynamic parameters calculated at 25 degrees C are as follows: DeltaG++ = 46.23 kJ/mol, DeltaH++ = 10.79 kJ/mol, and DeltaS++ = -119.72 J.mol(-1).K(-1) for Co(II)-loaded PfMetAP; DeltaG++ = 46.44 kJ/mol, DeltaH++ = 16.94 kJ/mol, and DeltaS++ = -99.67 J.mol(-1).K(-1) for Fe(II)-loaded PfMetAP. Interestingly, at higher temperatures (> 50 degrees C), Fe(II)-loaded PfMetAP-II is more active (1.4-fold at 85 degrees C) than Co(II)-loaded PfMetAP-II.  相似文献   

8.
V M D'souza  R C Holz 《Biochemistry》1999,38(34):11079-11085
The identity of the physiologically relevant metal ions for the methionyl aminopeptidase (MetAP) from Escherichia coli was investigated and is suggested to be Fe(II). The metal content of whole cells in the absence and presence of expression of the type I MetAP from E. coli was determined by inductively coupled plasma (ICP) emission analysis. The observed change in whole cell concentrations of cobalt, cadmium, copper, nickel, strontium, titanium, and vanadium upon expression of MetAP was negligible. On the other hand, significant increases in the cellular metal ion concentrations of chromium, zinc, manganese, and iron were observed with the increase in iron concentration being 4.4 and 6.2 times greater than that of manganese and zinc, respectively. Activity assays of freshly lysed BL21(DE3) cells containing the pMetAAP plasmid revealed detectable levels (>2 units/mg) of MetAP activity. Control experiments with BL21(DE3) without the MetAP plasmid showed no detectable enzymatic activity. Since MetAP is active upon expression, these data strongly suggest that cobalt is not the in vivo metal ion for the MetAP from E. coli. The MetAP from E. coli as purified was found to be catalytically inactive (相似文献   

9.
Ionizing radiation causes formation of thymine hydroperoxides in DNA. Their decomposition generates more stable products and active oxygen species which may oxidize other DNA bases. We have determined the effects of free and chelated metal ions on the degradation of 5-hydroperoxymethyl-2'-deoxyuridine (HPMdU). Two products were formed as analyzed by HPLC: 5-hydroxymethyl-2'-deoxyuridine (HMdU) and 5-formyl-2'-deoxyuridine (FdU). Sn(II) and Fe(II) caused instantaneous HPMdU degradation; Sn(II) generated only HMdU, whereas Fe(II) formed about equal amounts of both. Sn(IV) and Fe(III) were inactive. Cu(I), Cu(II), and Co(II) caused a time-dependent formation of both products, with FdU predominating. In the presence of Cu(I), Cu(II), and Fe(II), formate inhibited formation of HMdU but enhanced that of FdU. EDTA abolished Cu(I)-induced decomposition of HPMdU but only decreased that which was mediated by Cu(II). In contrast, EDTA enhanced the activity of Fe(III) with a time-dependent formation of FdU. EDTA and diethylenetriaminepentaacetic acid (DTPA) caused an instantaneous Fe(II)-mediated decomposition of HPMdU to FdU. Only desferal partially inhibited the activity of Fe(II), whereas the activities of Cu(I), Cu(II), and Fe(III) were blocked by desferal and DTPA. Possible mechanisms of HPMdU degradation by metal ions in the absence or presence of formate or chelators as well as formation of the .OH are discussed.  相似文献   

10.
11.
The gene encoding melamine deaminase (TriA) from Pseudomonas sp. strain NRRL B-12227 was identified, cloned into Escherichia coli, sequenced, and expressed for in vitro study of enzyme activity. Melamine deaminase displaced two of the three amino groups from melamine, producing ammeline and ammelide as sequential products. The first deamination reaction occurred more than 10 times faster than the second. Ammelide did not inhibit the first or second deamination reaction, suggesting that the lower rate of ammeline hydrolysis was due to differential substrate turnover rather than product inhibition. Remarkably, melamine deaminase is 98% identical to the enzyme atrazine chlorohydrolase (AtzA) from Pseudomonas sp. strain ADP. Each enzyme consists of 475 amino acids and differs by only 9 amino acids. AtzA was shown to exclusively catalyze dehalogenation of halo-substituted triazine ring compounds and had no activity with melamine and ammeline. Similarly, melamine deaminase had no detectable activity with the halo-triazine substrates. Melamine deaminase was active in deamination of a substrate that was structurally identical to atrazine, except for the substitution of an amino group for the chlorine atom. Moreover, melamine deaminase and AtzA are found in bacteria that grow on melamine and atrazine compounds, respectively. These data strongly suggest that the 9 amino acid differences between melamine deaminase and AtzA represent a short evolutionary pathway connecting enzymes catalyzing physiologically relevant deamination and dehalogenation reactions, respectively.  相似文献   

12.
The TrzN protein, which is involved in s-triazine herbicide catabolism by Arthrobacter aurescens TC1, was cloned and expressed in Escherichia coli as a His-tagged protein. The recombinant protein was purified via nickel column chromatography. The purified TrzN protein was tested with 31 s-triazine and pyrimidine ring compounds; 22 of the tested compounds were substrates. TrzN showed high activity with sulfur-substituted s-triazines and the highest activity with ametryn sulfoxide. Hydrolysis of ametryn sulfoxide by TrzN, both in vitro and in vivo, yielded a product(s) that reacted with 7-chloro-4-nitrobenz-2-oxa-1,3-diazole (NBD-Cl) to generate a diagnostic blue product. Atrazine chlorohydrolase, AtzA, did not hydrolyze ametryn sulfoxide, and no color was formed by amending those enzyme incubations with NBD-Cl. TrzN and AtzA could also be distinguished by reaction with ametryn. TrzN, but not AtzA, hydrolyzed ametryn to methylmercaptan. Methylmercaptan reacted with NBD-Cl to produce a diagnostic yellow product having an absorption maximum at 420 nm. The yellow color with ametryn was shown to selectively demonstrate the presence of TrzN, but not AtzA or other enzymes, in whole microbial cells. The present study was the first to purify an active TrzN protein in recombinant form and develop a colorimetric test for determining TrzN activity, and it significantly extends the known substrate range for TrzN.  相似文献   

13.
beta-Amyloid protein is the major component of senile plaques found in the brains of Alzheimer's patients. Previously, a new biochemical property of amyloid, its ability to disrupt ester and peptide bonds, was described [Elbaum, D., Brzyska, M., Bacia, A. & Alkon, D. (2000) Biochem. Biophys. Res. Commun. 267, 733-738]. In the present work we compare the ability of beta-amyloid to hydrolyse and oxidize model fluorescent derivatives of dichlorofluorescein [dichlorodihydrofluorescein (H2DCF) or dichlorofluorescein diacetate (DCF-DA), respectively] to the same final product (dichlorofluorescein). Although there is accumulating evidence of oxidative properties of beta-amyloid, little is known about its hydrolytic abilities. Chemical modification studies revealed that hydrolytic properties are related to a His, Ser and Asp/Glu triad, while residues of His, Tyr and Met are involved in the oxidative activity of amyloid. Studies with the rat homologue of human beta-amyloid (1-40), containing three amino-acid substitutions (Arg5-->Gly, Tyr10-->Phe and His13-->Arg) confirmed a role of His in the studied processes. Reduction of the hydrolysis product caused by inhibitors of Ser esterases (phenylmethylsulphonyl fluoride and eserine) suggests that beta-amyloid-mediated hydrolysis is Ser sensitive. Antioxidants and metal chelators that reduced H2DCF oxidation did not change or increase DCF-DA hydrolysis. Solvent isotope effects suggest the involvement of hydrogen bonds in the hydrolysis reaction. Hydrolysis was inhibited by redox-active metal ions and was practically oxygen independent while the oxidation process was redox-active-metal enhanced [Cu(II) and Fe(II) primarily], and oxygen dependent. Product formation was significantly inhibited by catalase and superoxide dismutase as well as benzoquinone, a specific superoxide anion radical scavenger. Increase of fluorescence by oxidation was strongly inhibited by azide and His and enhanced in samples prepared with deuterated phosphate buffer, suggesting singlet oxygen intermediacy. Our data are consistent with superoxide-mediated singlet oxygen intermediate in this Fenton mechanism-driven reaction. These results indicate that hydrolytic and oxidative properties of beta-amyloid are distinct features of this peptide and probably require different mechanisms to occur, but both of them may contribute to beta-amyloid toxicity.  相似文献   

14.
Complexes of Mn(II), Fe(III), Fe(II), Co(II), Ni(II), Cu(II), Zn(II) and Pt(II) with 2,6-diacetylpyridine bis(N4-azacyclic thiosemicarbazones), abbreviated as H2L, have been prepared and characterized by elemental analysis, molar conductance, magnetic moments (300-78 K) and spectral studies. On the basis of these studies, a distorted six-coordinate structure for Fe(L)Cl and a distorted five-coordinate structure for M(L) (M = Mn(II), Fe(II), Co(II), Ni(II), Cu(II), Zn(II), or Pt(II] are suggested. The ligands undergo deprotonation and appear to coordinate through the thione sulphur, the imine nitrogen and pyridyl nitrogen. All the ligands and metal complexes were screened for their antitumor activity against P 388 lymphocytic leukemia test system in mice, and it was found that a few of them possess significant activity at the dosages used.  相似文献   

15.
Catechol-O-methyltransferase (COMT) degrades catecholamines, such as dopamine and epinephrine, by methylating them in the presence of a divalent metal cation (usually Mg(II)), and S-adenosyl-L-methionine. The enzymatic activity of COMT is known to be vitally dependent on the nature of the bound metal: replacement of Mg(II) with Ca(II) leads to a complete deactivation of COMT; Fe(II) is slightly less than potent Mg(II), and Fe(III) is again an inhibitor. Considering the fairly modest role that the metal plays in the catalyzed reaction, this dependence is puzzling, and to date remains an enigma. Using a quantum mechanical / molecular mechanical dynamics method for extensive sampling of protein structure, and first principle quantum mechanical calculations for the subsequent mechanistic study, we explicate the effect of metal substitution on the rate determining step in the catalytic cycle of COMT, the methyl transfer. In full accord with experimental data, Mg(II) bound to COMT is the most potent of the studied cations and it is closely followed by Fe(II), whereas Fe(III) is unable to promote catalysis. In the case of Ca(II), a repacking of the protein binding site is observed, leading to a significant increase in the activation barrier and higher energy of reaction. Importantly, the origin of the effect of metal substitution is different for different metals: for Fe(III) it is the electronic effect, whereas in the case of Ca(II) it is instead the effect of suboptimal protein structure.  相似文献   

16.
In an effort to better understand the structure, metal content, the nature of the metal centers, and enzyme activity of Arabidopsis thaliana Glx2-2, the enzyme was overexpressed, purified, and characterized using metal analyses, kinetics, and UV–vis, EPR, and 1H NMR spectroscopies. Glx2-2-containing fractions that were purple, yellow, or colorless were separated during purification, and the differently colored fractions were found to contain different amounts of Fe and Zn(II). Spectroscopic analyses of the discrete fractions provided evidence for Fe(II), Fe(III), Fe(III)–Zn(II), and antiferromagnetically coupled Fe(II)–Fe(III) centers distributed among the discrete Glx2-2-containing fractions. The individual steady-state kinetic constants varied among the fractionated species, depending on the number and type of metal ion present. Intriguingly, however, the catalytic efficiency constant, k cat/K m, was invariant among the fractions. The value of k cat/K m governs the catalytic rate at low, physiological substrate concentrations. We suggest that the independence of k cat/K m on the precise makeup of the active-site metal center is evolutionarily related to the lack of selectivity for either Fe versus Zn(II) or Fe(II) versus Fe(III), in one or more metal binding sites.  相似文献   

17.
Adenine deaminase (ADE) catalyzes the conversion of adenine to hypoxanthine and ammonia. The enzyme isolated from Escherichia coli using standard expression conditions was low for the deamination of adenine (k(cat) = 2.0 s(-1); k(cat)/K(m) = 2.5 × 10(3) M(-1) s(-1)). However, when iron was sequestered with a metal chelator and the growth medium was supplemented with Mn(2+) prior to induction, the purified enzyme was substantially more active for the deamination of adenine with k(cat) and k(cat)/K(m) values of 200 s(-1) and 5 × 10(5) M(-1) s(-1), respectively. The apoenzyme was prepared and reconstituted with Fe(2+), Zn(2+), or Mn(2+). In each case, two enzyme equivalents of metal were necessary for reconstitution of the deaminase activity. This work provides the first example of any member of the deaminase subfamily of the amidohydrolase superfamily to utilize a binuclear metal center for the catalysis of a deamination reaction. [Fe(II)/Fe(II)]-ADE was oxidized to [Fe(III)/Fe(III)]-ADE with ferricyanide with inactivation of the deaminase activity. Reducing [Fe(III)/Fe(III)]-ADE with dithionite restored the deaminase activity, and thus, the diferrous form of the enzyme is essential for catalytic activity. No evidence of spin coupling between metal ions was evident by electron paramagnetic resonance or Mo?ssbauer spectroscopy. The three-dimensional structure of adenine deaminase from Agrobacterium tumefaciens (Atu4426) was determined by X-ray crystallography at 2.2 ? resolution, and adenine was modeled into the active site on the basis of homology to other members of the amidohydrolase superfamily. On the basis of the model of the adenine-ADE complex and subsequent mutagenesis experiments, the roles for each of the highly conserved residues were proposed. Solvent isotope effects, pH-rate profiles, and solvent viscosity were utilized to propose a chemical reaction mechanism and the identity of the rate-limiting steps.  相似文献   

18.
D'souza VM  Bennett B  Copik AJ  Holz RC 《Biochemistry》2000,39(13):3817-3826
The metal-binding properties of the methionyl aminopeptidase from Escherichia coli (MetAP) were investigated. Measurements of catalytic activity as a function of added Co(II) and Fe(II) revealed that maximal enzymatic activity is observed after the addition of only 1 equiv of divalent metal ion. Based on these studies, metal binding constants for the first metal binding event were found to be 0.3 +/- 0.2 microM and 0.2 +/- 0.2 microM for Co(II)- and Fe(II)-substituted MetAP, respectively. Binding of excess metal ions (>50 equiv) resulted in the loss of approximately 50% of the catalytic activity. Electronic absorption spectral titration of a 1 mM sample of MetAP with Co(II) provided a binding constant of 2.5 +/- 0.5 mM for the second metal binding site. Furthermore, the electronic absorption spectra of Co(II)-loaded MetAP indicated that both metal ions reside in a pentacoordinate geometry. Consistent with the absorption data, electron paramagnetic resonance (EPR) spectra of [CoCo(MetAP)] also indicated that the Co(II) geometries are not highly constrained, suggesting that each Co(II) ion in MetAP resides in a pentacoordinate geometry. EPR studies on [CoCo(MetAP)] also revealed that at pH 7.5 there is no significant spin-coupling between the two Co(II) ions, though a small proportion ( approximately 5%) of the sample exhibited detectable spin-spin interactions at pH values > 9.6. EPR studies on [Fe(III)_(MetAP)] and [Fe(III)Fe(III)(MetAP)] also suggested no spin-coupling between the two metal ions. (1)H nuclear magnetic resonance (NMR) spectra of [Co(II)_(MetAP)] in both H(2)O and D(2)O buffer indicated that the first metal binding site contains the only active-site histidine residue, His171. Mechanistic implications of the observed binding properties of divalent metal ions to the MetAP from E. coli are discussed.  相似文献   

19.
Stenchuk NN  Kapustiak KE 《Genetika》2003,39(8):1026-1032
A method of positive selection of mutants with impaired regulation of flavinogenesis and metal homeostasis in yeast Pichia guilliermondii was developed. This positive selection system was based on the isolation of pseudo-wild-type revertants (the Rib+ phenotype) in riboflavin-dependent rib1-86 mutant (the Rib- phenotype) of yeast P. guilliermondii. Mutation rib1-86 blocks activity of the GTP cyclohydrolase II catalyzing the first step in riboflavin (RF) biosynthesis. Study of a collection of spontaneous Rib+ revertants allowed the identification of a considerably large number of genetic loci responsible for the suppression of rib1-86, which include both previously identified three loci (rib80, rib81, and hit1) and six new loci designated red1-red6 (reduction). A comparative analysis of the wild-type strain and red mutants revealed that these mutants had higher activity levels of GTP cyclohydrolase and RF-synthase, elevated levels of RF biosynthesis, enhanced Fe/Cu reductase activity and higher total iron content in cells and that they are characterized by enhanced sensitivity to transition metals (Fe(III), Cu(II), Cd(II), Co(II), Zn(II), Ag(I), and to H2O2. The metal hypersensitivity of mutant cells can be prevented by an increased amount of extracellular iron ions. Mutations red1 and red6 synergistically interact with the locus rib81 in the course of RF biosynthesis. Obviously, each RED gene plays an important role in the regulation of both flavinogenesis and metal homeostasis in P. guilliermondii cells.  相似文献   

20.
In an effort to probe whether the metal content of metallo-beta-lactamase L1 is affected by metal ion bioavailability, L1 was overexpressed as mature protein (M-L1) and full-length (FL-L1) analogues, and the analogues were characterized with metal analyses, kinetics, and EPR spectroscopy. FL-L1, containing the putative leader sequence, was localized in the periplasm of Escherichia coli and shown to bind Zn(II) preferentially. The metal content of FL-L1 could be altered if the enzyme was overexpressed in minimal medium containing Fe and Mn, and surprisingly, an Fe-binding analogue was obtained. On the other hand, M-L1, lacking the putative leader sequence, was localized in the cytoplasm of E. coli and shown to bind various amounts of Fe and Zn(II), and like FL-L1, the metal content of the resulting enzyme could be affected by the amount of metal ions in the growth medium. L1 was refolded in the presence of Fe, and a dinuclear Fe-containing analogue of L1 was obtained, although this analogue is catalytically inactive. EPR spectra demonstrate the presence of an antiferromagnetically coupled Fe(III)Fe(II) center in Fe-containing L1 and suggest the presence of a Fe(III)Zn(II) center in M-L1. Metal analyses on the cytoplasmic and periplasmic fractions of E. coli showed that the concentration of metal ions in the periplasm is not tightly controlled and increases as the concentration of metal ions in the growth medium increases. In contrast, the concentration of Zn(II) in the cytoplasm is tightly controlled while that of Fe is less so.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号